23
Published as a conference paper at ICLR 2018 MEASURING THE I NTRINSIC D IMENSION OF O BJECTIVE L ANDSCAPES Chunyuan Li * Duke University [email protected] Heerad Farkhoor, Rosanne Liu, and Jason Yosinski Uber AI Labs {heerad,rosanne,yosinski}@uber.com ABSTRACT Many recently trained neural networks employ large numbers of parameters to achieve good performance. One may intuitively use the number of parameters required as a rough gauge of the difficulty of a problem. But how accurate are such notions? How many parameters are really needed? In this paper we at- tempt to answer this question by training networks not in their native parameter space, but instead in a smaller, randomly oriented subspace. We slowly increase the dimension of this subspace, note at which dimension solutions first appear, and define this to be the intrinsic dimension of the objective landscape. The ap- proach is simple to implement, computationally tractable, and produces several suggestive conclusions. Many problems have smaller intrinsic dimensions than one might suspect, and the intrinsic dimension for a given dataset varies little across a family of models with vastly different sizes. This latter result has the profound implication that once a parameter space is large enough to solve a prob- lem, extra parameters serve directly to increase the dimensionality of the solution manifold. Intrinsic dimension allows some quantitative comparison of problem difficulty across supervised, reinforcement, and other types of learning where we conclude, for example, that solving the inverted pendulum problem is 100 times easier than classifying digits from MNIST, and playing Atari Pong from pixels is about as hard as classifying CIFAR-10. In addition to providing new cartogra- phy of the objective landscapes wandered by parameterized models, the method is a simple technique for constructively obtaining an upper bound on the mini- mum description length of a solution. A byproduct of this construction is a simple approach for compressing networks, in some cases by more than 100 times. 1 I NTRODUCTION Training a neural network to model a given dataset entails several steps. First, the network designer chooses a loss function and a network architecture for a given dataset. The architecture is then ini- tialized by populating its weights with random values drawn from some distribution. Finally, the network is trained by adjusting its weights to produce a loss as low as possible. We can think of the training procedure as traversing some path along an objective landscape. Note that as soon as a dataset and network architecture are specified, the landscape in its entirety is completely deter- mined. It is instantiated and frozen; all subsequent parameter initialization, forward and backward propagation, and gradient steps taken by an optimizer are just details of how the frozen space is explored. Consider a network parameterized by D weights. We can picture its associated objective landscape as a set of “hills and valleys” in D dimensions, where each point in R D corresponds to a value of the loss, i.e., the elevation of the landscape. If D =2, the map from two coordinates to one scalar loss can be easily imagined and intuitively understood by those living in a three-dimensional world with similar hills. However, in higher dimensions, our intuitions may not be so faithful, and generally we must be careful, as extrapolating low-dimensional intuitions to higher dimensions can lead to un- reliable conclusions. The difficulty of understanding high-dimensional landscapes notwithstanding, it is the lot of neural network researchers to spend their efforts leading (or following?) networks * Work performed as an intern at Uber AI Labs. 1 arXiv:1804.08838v1 [cs.LG] 24 Apr 2018

A arXiv:1804.08838v1 [cs.LG] 24 Apr 2018 · fheerad,rosanne,[email protected] ... serve directly to increase the dimensionality of the solution ... parameter vectors for direct optimization

Embed Size (px)

Citation preview

Published as a conference paper at ICLR 2018

MEASURING THE INTRINSIC DIMENSIONOF OBJECTIVE LANDSCAPES

Chunyuan Li ∗Duke [email protected]

Heerad Farkhoor, Rosanne Liu, and Jason YosinskiUber AI Labs{heerad,rosanne,yosinski}@uber.com

ABSTRACT

Many recently trained neural networks employ large numbers of parameters toachieve good performance. One may intuitively use the number of parametersrequired as a rough gauge of the difficulty of a problem. But how accurate aresuch notions? How many parameters are really needed? In this paper we at-tempt to answer this question by training networks not in their native parameterspace, but instead in a smaller, randomly oriented subspace. We slowly increasethe dimension of this subspace, note at which dimension solutions first appear,and define this to be the intrinsic dimension of the objective landscape. The ap-proach is simple to implement, computationally tractable, and produces severalsuggestive conclusions. Many problems have smaller intrinsic dimensions thanone might suspect, and the intrinsic dimension for a given dataset varies littleacross a family of models with vastly different sizes. This latter result has theprofound implication that once a parameter space is large enough to solve a prob-lem, extra parameters serve directly to increase the dimensionality of the solutionmanifold. Intrinsic dimension allows some quantitative comparison of problemdifficulty across supervised, reinforcement, and other types of learning where weconclude, for example, that solving the inverted pendulum problem is 100 timeseasier than classifying digits from MNIST, and playing Atari Pong from pixelsis about as hard as classifying CIFAR-10. In addition to providing new cartogra-phy of the objective landscapes wandered by parameterized models, the methodis a simple technique for constructively obtaining an upper bound on the mini-mum description length of a solution. A byproduct of this construction is a simpleapproach for compressing networks, in some cases by more than 100 times.

1 INTRODUCTION

Training a neural network to model a given dataset entails several steps. First, the network designerchooses a loss function and a network architecture for a given dataset. The architecture is then ini-tialized by populating its weights with random values drawn from some distribution. Finally, thenetwork is trained by adjusting its weights to produce a loss as low as possible. We can think ofthe training procedure as traversing some path along an objective landscape. Note that as soon asa dataset and network architecture are specified, the landscape in its entirety is completely deter-mined. It is instantiated and frozen; all subsequent parameter initialization, forward and backwardpropagation, and gradient steps taken by an optimizer are just details of how the frozen space isexplored.

Consider a network parameterized by D weights. We can picture its associated objective landscapeas a set of “hills and valleys” inD dimensions, where each point in RD corresponds to a value of theloss, i.e., the elevation of the landscape. If D = 2, the map from two coordinates to one scalar losscan be easily imagined and intuitively understood by those living in a three-dimensional world withsimilar hills. However, in higher dimensions, our intuitions may not be so faithful, and generally wemust be careful, as extrapolating low-dimensional intuitions to higher dimensions can lead to un-reliable conclusions. The difficulty of understanding high-dimensional landscapes notwithstanding,it is the lot of neural network researchers to spend their efforts leading (or following?) networks

∗Work performed as an intern at Uber AI Labs.

1

arX

iv:1

804.

0883

8v1

[cs

.LG

] 2

4 A

pr 2

018

Published as a conference paper at ICLR 2018

over these multi-dimensional surfaces. Therefore, any interpreted geography of these landscapes isvaluable.

Several papers have shed valuable light on this landscape, particularly by pointing out flaws in com-mon extrapolation from low-dimensional reasoning. Dauphin et al. (2014) showed that, in contrastto conventional thinking about getting stuck in local optima (as one might be stuck in a valley inour familiar D = 2), local critical points in high dimension are almost never valleys but are insteadsaddlepoints: structures which are “valleys” along a multitude of dimensions with “exits” in a mul-titude of other dimensions. The striking conclusion is that one has less to fear becoming hemmed inon all sides by higher loss but more to fear being waylaid nearly indefinitely by nearly flat regions.Goodfellow et al. (2015) showed another property: that paths directly from the initial point to thefinal point of optimization are often monotonically decreasing. Though dimension is high, the spaceis in some sense simpler than we thought: rather than winding around hills and through long twistingcorridors, the walk could just as well have taken a straight line without encountering any obstacles,if only the direction of the line could have been determined at the outset.

In this paper we seek further understanding of the structure of the objective landscape by restrictingtraining to random slices through it, allowing optimization to proceed in randomly generated sub-spaces of the full parameter space. Whereas standard neural network training involves computinga gradient and taking a step in the full parameter space (RD above), we instead choose a randomd-dimensional subspace of RD, where generally d < D, and optimize directly in this subspace. Byperforming experiments with gradually larger values of d, we can find the subspace dimension atwhich solutions first appear, which we call the measured intrinsic dimension of a particular problem.Examining intrinsic dimensions across a variety of problems leads to a few new intuitions about theoptimization problems that arise from neural network models.

We begin in Sec. 2 by defining more precisely the notion of intrinsic dimension as a measure ofthe difficulty of objective landscapes. In Sec. 3 we measure intrinsic dimension over a variety ofnetwork types and datasets, including MNIST, CIFAR-10, ImageNet, and several RL tasks. Basedon these measurements, we draw a few insights on network behavior, and we conclude in Sec. 4.

2 DEFINING AND ESTIMATING INTRINSIC DIMENSION

We introduce the intrinsic dimension of an objective landscape with an illustrative toy problem.Let θ(D) ∈ RD be a parameter vector in a parameter space of dimension D, let θ(D)

0 be a randomlychosen initial parameter vector, and let θ(D)

∗ be the final parameter vector arrived at via optimization.

Consider a toy optimization problem where D = 1000 and where θ(D) optimized to minimize asquared error cost function that requires the first 100 elements to sum to 1, the second 100 elementsto sum to 2, and so on until the vector has been divided into 10 groups with their requisite 10 sums.We may start from a θ(D)

0 that is drawn from a Gaussian distribution and optimize in RD to find aθ(D)∗ that solves the problem with cost arbitrarily close to zero.

Solutions to this problem are highly redundant. With a little algebra, one can find that the manifoldof solutions is a 990 dimensional hyperplane: from any point that has zero cost, there are 990orthogonal directions one can move and remain at zero cost. Denoting as s the dimensionality ofthe solution set, we define the intrinsic dimensionality dint of a solution as the codimension of thesolution set inside of RD:

D = dint + s (1)

Here the intrinsic dimension dint is 10 (1000 = 10 + 990), with 10 corresponding intuitively to thenumber of constraints placed on the parameter vector. Though the space is large (D = 1000), thenumber of things one needs to get right is small (dint = 10).

2.1 MEASURING INTRINSIC DIMENSION VIA RANDOM SUBSPACE TRAINING

The above example had a simple enough form that we obtained dint = 10 by calculation. Butin general we desire a method to measure or approximate dint for more complicated problems,including problems with data-dependent objective functions, e.g. neural network training. Randomsubspace optimization provides such a method.

2

Published as a conference paper at ICLR 2018

0 10 20 30 40 50Subspace dim d

0.00

0.25

0.50

0.75

1.00

Per

form

ance

dint90 = dint100 = 10

Figure 1: (left) Illustration of parameter vectors for direct optimization in the D = 3 case. (middle)Illustration of parameter vectors and a possible random subspace for the D = 3, d = 2 case. (right)Plot of performance vs. subspace dimension for the toy example of toy example of Sec. 2. Theproblem becomes both 90% solvable and 100% solvable at random subspace dimension 10, sodint90 and dint100 are 10.

Standard optimization, which we will refer to hereafter as the direct method of training, entailsevaluating the gradient of a loss with respect to θ(D) and taking steps directly in the space of θ(D).To train in a random subspace, we instead define θ(D) in the following way:

θ(D) = θ(D)0 + Pθ(d) (2)

where P is a randomly generated D × d projection matrix1 and θ(d) is a parameter vector in a gen-erally smaller space Rd. θ(D)

0 and P are randomly generated and frozen (not trained), so the systemhas only d degrees of freedom. We initialize θ(d) to a vector of all zeros, so initially θ(D) = θ

(D)0 .

This convention serves an important purpose for neural network training: it allows the network tobenefit from beginning in a region of parameter space designed by any number of good initializationschemes (Glorot & Bengio, 2010; He et al., 2015) to be well-conditioned, such that gradient descentvia commonly used optimizers will tend to work well.2

Training proceeds by computing gradients with respect to θ(d) and taking steps in that space.Columns of P are normalized to unit length, so steps of unit length in θ(d) chart out unit lengthmotions of θ(D). Columns of P may also be orthogonalized if desired, but in our experimentswe relied simply on the approximate orthogonality of high dimensional random vectors. By thisconstruction P forms an approximately orthonormal basis for a randomly oriented d dimensionalsubspace of RD, with the origin of the new coordinate system at θ(D)

0 . Fig. 1 (left and middle) showsan illustration of the related vectors.

Consider a few properties of this training approach. If d = D and P is a large identity matrix, werecover exactly the direct optimization problem. If d = D but P is instead a random orthonormalbasis for all of RD (just a random rotation matrix), we recover a rotated version of the direct problem.Note that for some “rotation-invariant” optimizers, such as SGD and SGD with momentum, rotatingthe basis will not change the steps taken nor the solution found, but for optimizers with axis-alignedassumptions, such as RMSProp (Tieleman & Hinton, 2012) and Adam (Kingma & Ba, 2014), thepath taken through θ(D) space by an optimizer will depend on the rotation chosen. Finally, in thegeneral case where d < D and solutions exist in D, solutions will almost surely (with probability 1)not be found if d is less than the codimension of the solution. On the other hand, when d ≥ D−s, ifthe solution set is a hyperplane, the solution will almost surely intersect the subspace, but for solutionsets of arbitrary topology, intersection is not guaranteed. Nonetheless, by iteratively increasing d,re-running optimization, and checking for solutions, we obtain one estimate of dint. We try thissweep of d for our toy problem laid out in the beginning of this section, measuring (by convention

1This projection matrix can take a variety of forms each with different computational considerations. Inlater sections we consider dense, sparse, and implicit P matrices.

2A second, more subtle reason to start away from the origin and with a randomly oriented subspace isthat this puts the subspace used for training and the solution set in general position with respect to each other.Intuitively, this avoids pathological cases where both the solution set and the random subspace contain structureoriented around the origin or along axes, which could bias toward non-intersection or toward intersection,depending on the problem.

3

Published as a conference paper at ICLR 2018

as described in the next section) the positive performance (higher is better) instead of loss.3 Asexpected, the solutions are first found at d = 10 (see Fig. 1, right), confirming our intuition that forthis problem, dint = 10.

2.2 DETAILS AND CONVENTIONS

In the rest of this paper, we measure intrinsic dimensions for particular neural network problemsand draw conclusions about the associated objective landscapes and solution sets. Because model-ing real data is more complex than the above toy example, and losses are generally never exactlyzero, we first choose a heuristic for classifying points on the objective landscape as solutions vs.non-solutions. The heuristic we choose is to threshold network performance at some level rela-tive to a baseline model, where generally we take as baseline the best directly trained model. Insupervised classification settings, validation accuracy is used as the measure of performance, andin reinforcement learning scenarios, the total reward (shifted up or down such that the minimumreward is 0) is used. Accuracy and reward are preferred to loss to ensure results are grounded toreal-world performance and to allow comparison across models with differing scales of loss anddifferent amounts of regularization included in the loss.

We define dint100 as the intrinsic dimension of the “100%” solution: solutions whose performanceis statistically indistinguishable from baseline solutions. However, when attempting to measuredint100, we observed it to vary widely, for a few confounding reasons: dint100 can be very high— nearly as high as D — when the task requires matching a very well-tuned baseline model, butcan drop significantly when the regularization effect of restricting parameters to a subspace boostsperformance by tiny amounts. While these are interesting effects, we primarily set out to measurethe basic difficulty of problems and the degrees of freedom needed to solve (or approximately solve)them rather than these subtler effects.

Thus, we found it more practical and useful to define and measure dint90 as the intrinsic dimensionof the “90%” solution: solutions with performance at least 90% of the baseline.

We chose 90% after looking at a number of dimension vs. performance plots (e.g. Fig. 2) as areasonable trade off between wanting to guarantee solutions are as good as possible, but also wantingmeasured dint values to be robust to small noise in measured performance. If too high a thresholdis used, then the dimension at which performance crosses the threshold changes a lot for only tinychanges in accuracy, and we always observe tiny changes in accuracy due to training noise.

If a somewhat different (higher or lower) threshold were chosen, we expect most of conclusions inthe rest of the paper to remain qualitatively unchanged. In the future, researchers may find it usefulto measure dint using higher or lower thresholds.

3 RESULTS AND DISCUSSION

3.1 MNIST

We begin by analyzing a fully connected (FC) classifier trained on MNIST. We choose a networkwith layer sizes 784–200–200–10, i.e. a network with two hidden layers of width 200; this resultsin a total number of parameters D = 199, 210. A series of experiments with gradually increasingsubspace dimension d produce monotonically increasing performances, as shown in Fig. 2 (left).By checking the subspace dimension at which performance crosses the 90% mark, we measure thisnetwork’s intrinsic dimension dint90 at about 750.

Some networks are very compressible. A salient initial conclusion is that 750 is quite low. Atthat subspace dimension, only 750 degrees of freedom (0.4%) are being used and 198,460 (99.6%)unused to obtain 90% of the performance of the direct baseline model. A compelling corollary ofthis result is a simple, new way of creating and training compressed networks, particularly networksfor applications in which the absolute best performance is not critical. To store this network, oneneed only store a tuple of three items: (i) the random seed to generate the frozen θ(D)

0 , (ii) the

3For this toy problem we define performance = exp(−loss), bounding performance between 0 and 1,with 1 being a perfect solution.

4

Published as a conference paper at ICLR 2018

0 200 400 600 800 1000 1200 1400Subspace dim d

0.2

0.4

0.6

0.8

1.0

Val

idat

ion

accu

racy

baseline

90% baseline

0 200 400 600 800 1000Subspace dim d

0.2

0.4

0.6

0.8

1.0

Val

idat

ion

accu

racy

baseline

90% baseline

Figure 2: Performance (validation accuracy) vs. subspace dimension d for two networks trainedon MNIST: (left) a 784–200–200–10 fully-connected (FC) network (D = 199,210) and (right) aconvolutional network, LeNet (D = 44,426). The solid line shows performance of a well-traineddirect (FC or conv) model, and the dashed line shows the 90% threshold we use to define dint90. Thestandard derivation of validation accuracy and measured dint90 are visualized as the blue vertical andred horizontal error bars. We oversample the region around the threshold to estimate the dimensionof crossing more exactly. We use one-run measurements for dint90 of 750 and 290, respectively.

random seed to generate P and (iii) the 750 floating point numbers in θ(d)∗ . It leads to compression(assuming 32-bit floats) by a factor of 260× from 793kB to only 3.2kB, or 0.4% of the full parametersize. Such compression could be very useful for scenarios where storage or bandwidth are limited,e.g. including neural networks in downloaded mobile apps or on web pages.

This compression approach differs from other neural network compression methods in the followingaspects. (i) While it has previously been appreciated that large networks waste parameters (Dauphin& Bengio, 2013) and weights contain redundancy (Denil et al., 2013) that can be exploited for post-hoc compression (Wen et al., 2016), this paper’s method constitutes a much simpler approach tocompression, where training happens once, end-to-end, and where any parameterized model is anallowable base model. (ii) Unlike layerwise compression models (Denil et al., 2013; Wen et al.,2016), we operate in the entire parameter space, which could work better or worse, depending onthe network. (iii) Compared to methods like that of Louizos et al. (2017), who take a Bayesianperspective and consider redundancy on the level of groups of parameters (input weights to a singleneuron) by using group-sparsity-inducing hierarchical priors on the weights, our approach is simplerbut not likely to lead to compression as high as the levels they attain. (iv) Our approach only reducesthe number of degrees of freedom, not the number of bits required to store each degree of freedom,e.g. as could be accomplished by quantizing weights (Han et al., 2016). Both approaches couldbe combined. (v) There is a beautiful array of papers on compressing networks such that they alsoachieve computational savings during the forward pass (Wen et al., 2016; Han et al., 2016; Yanget al., 2015); subspace training does not speed up execution time during inference. (vi) Finally, notethe relationships between weight pruning, weight tying, and subspace training: weight pruning isequivalent to finding, post-hoc, a subspace that is orthogonal to certain axes of the full parameterspace and that intersects those axes at the origin. Weight tying, e.g. by random hashing of weightsinto buckets (Chen et al., 2015), is equivalent to subspace training where the subspace is restrictedto lie along the equidistant “diagonals” between any axes that are tied together.

Robustness of intrinsic dimension. Next, we investigate how intrinsic dimension varies acrossFC networks with a varying number of layers and varying layer width.4 We perform a grid sweepof networks with number of hidden layers L chosen from {1, 2, 3, 4, 5} and width W chosen from{50, 100, 200, 400}. Fig. S6 in the Supplementary Information shows performance vs. subspacedimension plots in the style of Fig. 2 for all 20 networks, and Fig. 3 shows each network’s dint90plotted against its native dimension D. As one can see, D changes by a factor of 24.1 between thesmallest and largest networks, but dint90 changes over this range by a factor of only 1.33, with muchof this possibly due to noise.

Thus it turns out that the intrinsic dimension changes little even as models grown in width or depth!The striking conclusion is that every extra parameter added to the network — every extra dimensionadded to D — just ends up adding one dimension to the redundancy of the solution, s.

4Note that here we used a global baseline of 100% accuracy to compare simply and fairly across all models.See Sec. S5 for similar results obtained using instead 20 separate baselines for each of the 20 models.

5

Published as a conference paper at ICLR 2018

Figure 3: Measured intrinsic dimension dint90 vs number of parameters D for 20 FC models ofvarying width (from 50 to 400) and depth (number of hidden layers from 1 to 5) trained on MNIST.The red interval is the standard derivation of the measurement of dint90. Though the number ofnative parameters D varies by a factor of 24.1, dint90 varies by only 1.33, with much of that factorpossibly due to noise, showing that dint90 is a fairly robust measure across a model family and thateach extra parameter ends up adding an extra dimension directly to the redundancy of the solution.Standard deviation was estimated via bootstrap; see Sec. S5.1.

101 102 103 104

Number of trainable parameters

0.2

0.4

0.6

0.8

1.0

Val

idat

ion

accu

racy

Direct

Subspace

102 103 104

Number of trainable parameters

0.2

0.4

0.6

0.8

1.0V

alid

atio

nac

cura

cy

Direct

Subspace

Figure 4: Performance vs. number of trainable parameters for (left) FC networks and (right) convo-lutional networks trained on MNIST. Randomly generated direct networks are shown (gray circles)alongside all random subspace training results (blue circles) from the sweep shown in Fig. S6. FCnetworks show a persistent gap in dimension, suggesting general parameter inefficiency of FC mod-els. The parameter efficiency of convolutional networks varies, as the gray points can be significantlyto the right of or close to the blue manifold.

Often the most accurate directly trained models for a problem have far more parameters than needed(Zhang et al., 2017); this may be because they are just easier to train, and our observation suggestsa reason why: with larger models, solutions have greater redundancy and in a sense “cover” moreof the space.5 To our knowledge, this is the first time this phenomenon has been directly measured.We should also be careful not to claim that all FC nets on MNIST will have an intrinsic dimensionof around 750; instead, we should just consider that we have found for this architecture/datasetcombination a wide plateau of hyperparamter space over which intrinsic dimension is approximatelyconstant.

Are random subspaces really more parameter-efficient for FC nets? One might wonder towhat extent claiming 750 parameters is meaningful given that performance achieved (90%) is farworse than a state of the art network trained on MNIST. With such a low bar for performance, coulda directly trained network with a comparable number of trainable parameters be found that achievesthe same performance? We generated 1000 small networks (depth randomly chosen from {1, 2, 3,4, 5}, layer width randomly from {2, 3, 5, 8, 10, 15, 20, 25}, seed set randomly) in an attempt to

5To be precise, we may not conclude “greater coverage” in terms of the volume of the solution set —volumes are not comparable across spaces of different dimension, and our measurements have only estimatedthe dimension of the solution set, not its volume. A conclusion we may make is that as extra parameters areadded, the ratio of solution dimension to total dimension, s/D, increases, approaching 1. Further researchcould address other notions of coverage.

6

Published as a conference paper at ICLR 2018

find high-performing, small FC networks, but as Fig. 4 (left) shows, a gap still exists between thesubspace dimension and the smallest direct FC network giving the same performance at most levelsof performance.

Measuring dint90 on a convolutional network. Next we measure dint90 of a convolutional net-work, LeNet (D=44,426). Fig. 2 (right) shows validation accuracy vs. subspace dimension d, andwe find dint90 = 290, or a compression rate of about 150× for this network. As with the FC caseabove, we also do a sweep of random networks,

but notice that the performance gap of convnets between direct and subspace training methods be-comes closer for fixed budgets, i.e., the number of trainable parameters. Further, the performanceof direct training varies significantly, depending on the extrinsic design of convet architectures. Weinterpret these results in terms of the Minimum Description Length below.

Relationship between Intrinsic Dimension and Minimum Description Length (MDL). As dis-cussed earlier, the random subspace training method leads naturally to a compressed representationof a network, where only d floating point numbers need to be stored. We can consider this d as anupper bound on the MDL of the problem solution.6 We cannot yet conclude the extent to which thisbound is loose or tight, and tightness may vary by problem. However, to the extent that it is tighterthan previous bounds (e.g., just the number of parameters D) and to the extent that it is correlatedwith the actual MDL, we can use this interpretation to judge which solutions are more well-suitedto the problem in a principled way. As developed by Rissanen (1978) and further by Hinton &Van Camp (1993), holding accuracy constant, the best model is the one with the shortest MDL.

Thus, there is some rigor behind our intuitive assumption that LeNet is a better model than an FCnetwork for MNIST image classification, because its intrinsic dimension is lower (dint90 of 290 vs.750). In this particular case we are lead to a predictable conclusion, but as models become larger,more complex, and more heterogeneous, conclusions of this type will often not be obvious. Havinga simple method of approximating MDL may prove extremely useful for guiding model exploration,for example, for the countless datasets less well-studied than MNIST and for models consisting ofseparate sub-models that may be individually designed and evaluated (Ren et al., 2015; Kaiser et al.,2017). In this latter case, considering the MDL for a sub-model could provide a more detailedview of that sub-model’s properties than would be available by just analyzing the system’s overallvalidation performance.

Finally, note that although our approach is related to a rich body of work on estimating the “intrinsicdimension of a dataset” (Camastra & Vinciarelli, 2002; Kegl, 2003; Fukunaga & Olsen, 1971; Lev-ina & Bickel, 2005; Tenenbaum et al., 2000), it differs in a few respects. Here we do not measurethe number of degrees of freedom necessary to represent a dataset (which requires representationof a global p(X) and per-example properties and thus grows with the size of the dataset), but thoserequired to represent a model for part of the dataset (here p(y|X), which intuitively might saturateat some complexity even as a dataset grows very large). That said, in the following section we doshow measurements for a corner case where the model must memorize per-example properties.

Are convnets always better on MNIST? Measuring dint90 on shuffled data. Zhang et al. (2017)provocatively showed that large networks normally thought to generalize well can nearly as easily betrained to memorize entire training sets with randomly assigned labels or with input pixels providedin random order. Consider two identically sized networks: one trained on a real, non-shuffled datasetand another trained with shuffled pixels or labels. As noted by Zhang et al. (2017), externally thenetworks are very similar, and the training loss may even be identical at the final epoch. However,the intrinsic dimension of each may be measured to expose the differences in problem difficulty.When training on a dataset with shuffled pixels — pixels for each example in the dataset subjectto a random permutation, chosen once for the entire dataset — the intrinsic dimension of an FCnetwork remains the same at 750, because FC networks are invariant to input permutation. But theintrinsic dimension of a convnet increases from 290 to 1400, even higher than an FC network. Thuswhile convnets are better suited to classifying digits given images with local structure, when thisstructure is removed, violating convolutional assumptions, our measure can clearly reveal that many

6We consider MDL in terms of number of degrees of freedom instead of bits. For degrees of freedom storedwith constant fidelity (e.g. float32), these quantities are related by a constant factor (e.g. 32).

7

Published as a conference paper at ICLR 2018

Table 1: Measured dint90 on various supervised and reinforcement learning problems.

Dataset MNIST MNIST (Shuf Pixels) MNIST (Shuf Labels)Network Type FC LeNet FC LeNet FC

Parameter Dim. D 199,210 44,426 199,210 44,426 959,610Intrinsic Dim. dint90 750 290 750 1,400 190,000

... CIFAR-10 ImageNet Inverted Pendulum Humanoid Atari Pong

... FC LeNet SqueezeNet FC FC ConvNet

... 656,810 62,006 1,248,424 562 166,673 1,005,974

... 9,000 2,900 > 500k 4 700 6,000

more degrees of freedom are now required to model the underlying distribution. When training onMNIST with shuffled labels — the label for each example is randomly chosen — we redefine ourmeasure of dint90 relative to training accuracy (validation accuracy is always at chance). We findthat memorizing random labels on the 50,000 example MNIST training set requires a very highdimension, dint90 = 190, 000, or 3.8 floats per memorized label. Sec. S5.2 gives a few furtherresults, in particular that the more labels are memorized, the more efficient memorization is in termsof floats per label. Thus, while the network obviously does not generalize to an unseen validationset, it would seem “generalization” within a training set may be occurring as the network builds ashared infrastructure that makes it possible to more efficiently memorize labels.

3.2 CIFAR-10 AND IMAGENET

We scale to larger supervised classification problems by considering CIFAR-10 (Krizhevsky & Hin-ton, 2009) and ImageNet (Russakovsky et al., 2015). When scaling beyond MNIST-sized networkswith D on the order of 200k and d on the order of 1k, we find it necessary to use more efficientmethods of generating and projecting from random subspaces. This is particularly true in the caseof ImageNet, where the direct network can easily require millions of parameters. In Sec. S7, wedescribe and characterize scaling properties of three methods of projection: dense matrix projection,sparse matrix projection (Li et al., 2006), and the remarkable Fastfood transform (Le et al., 2013).We generally use the sparse projection method to train networks on CIFAR-10 and the Fastfoodtransform for ImageNet.

Measured dint90 values for CIFAR-10 and are ImageNet given in Table 1, next to all previousMNIST results and RL results to come. For CIFAR-10 we find qualitatively similar results toMNIST, but with generally higher dimension (9k vs. 750 for FC and 2.9k vs. 290 for LeNet).It is also interesting to observe the difference of dint90 across network architectures. For example,to achieve a global >50% validation accuracy on CIFAR-10, FC, LeNet and ResNet approximatelyrequires dint90 = 9k, 2.9k and 1k, respectively, showing that ResNets are more efficient. Full resultsand experiment details are given in Sec. S8 and Sec. S9. Due to limited time and memory issues,training on ImageNet has not yet given a reliable estimate for dint90 except that it is over 500k.

3.3 REINFORCEMENT LEARNING ENVIRONMENTS

Measuring intrinsic dimension allows us to perform some comparison across the divide betweensupervised learning and reinforcement learning. In this section we measure the intrinsic dimensionof three control tasks of varying difficulties using both value-based and policy-based algorithms.The value-based algorithm we evaluate is the Deep Q-Network (DQN) (Mnih et al., 2013), and thepolicy-based algorithm is Evolutionary Strategies (ES) (Salimans et al., 2017). Training details aregiven in Sec. S6.2. For all tasks, performance is defined as the maximum-attained (over trainingiterations) mean evaluation reward (averaged over 30 evaluations for a given parameter setting). InFig. 5, we show results of ES on three tasks: InvertedPendulum−v1, Humanoid−v1 in Mu-JoCo (Todorov et al., 2012), and Pong−v0 in Atari. Dots in each plot correspond to the (noisy)median of observed performance values across many runs for each given d, and the vertical uncer-tainty bar shows the maximum and minimum observed performance values. The dotted horizontalline corresponds to the usual 90% baseline derived from the best directly-trained network (the solid

8

Published as a conference paper at ICLR 2018

2 4 6 8 10Subspace dim d

0

200

400

600

800

1000

Best

eva

luat

ion

rewa

rd

100 200 300 400 500 600 700Subspace dim d

1000

2000

3000

4000

5000

6000

7000

8000

Best

eva

luat

ion

rewa

rd

2000 4000 6000 8000 10000

Subspace dim d

−20

−15

−10

−5

0

5

10

15

20

Bes

tev

alua

tion

rew

ard

Figure 5: Results using the policy-based ES algorithm to train agents on (left column)InvertedPendulum−v1, (middle column) Humanoid−v1, and (right column) Pong−v0. Theintrinsic dimensions found are 4, 700, and 6k. This places the walking humanoid task on a similarlevel of difficulty as modeling MNIST with a FC network (far less than modeling CIFAR-10 with aconvnet), and Pong on the same order of modeling CIFAR-10.

horizontal line). A dot is darkened signifying the first d that allows a satisfactory performance. Wefind that the inverted pendulum task is surprisingly easy, with dint100 = dint90 = 4, meaning thatonly four parameters are needed to perfectly solve the problem (see Stanley & Miikkulainen (2002)for a similarly small solution found via evolution). The walking humanoid task is more difficult: so-lutions are found reliably by dimension 700, a similar complexity to that required to model MNISTwith an FC network, and far less than modeling CIFAR-10 with a convnet. Finally, to play Pongon Atari (directly from pixels) requires a network trained in a 6k dimensional subspace, making iton the same order of modeling CIFAR-10. For an easy side-by-side comparison we list all intrinsicdimension values found for all problems in Table 1. For more complete ES results see Sec. S6.2,and Sec. S6.1 for DQN results.

4 CONCLUSIONS AND FUTURE DIRECTIONS

In this paper, we have defined the intrinsic dimension of objective landscapes and shown a simplemethod — random subspace training — of approximating it for neural network modeling problems.We use this approach to compare problem difficulty within and across domains. We find in somecases the intrinsic dimension is much lower than the direct parameter dimension, and hence enablenetwork compression, and in other cases the intrinsic dimension is similar to that of the best tunedmodels, and suggesting those models are better suited to the problem.

Further work could also identify better ways of creating subspaces for reparameterization: herewe chose random linear subspaces, but one might carefully construct other linear or non-linearsubspaces to be even more likely to contain solutions. Finally, as the field departs from single stack-of-layers image classification models toward larger and more heterogeneous networks (Ren et al.,2015; Kaiser et al., 2017) often composed of many modules and trained by many losses, methodslike measuring intrinsic dimension that allow some automatic assessment of model componentsmight provide much-needed greater understanding of individual black-box module properties.

ACKNOWLEDGMENTS

The authors gratefully acknowledge Zoubin Ghahramani, Peter Dayan, Sam Greydanus, Jeff Clune,and Ken Stanley for insightful discussions, Joel Lehman for initial idea validation, Felipe Such,Edoardo Conti and Xingwen Zhang for helping scale the ES experiments to the cluster, VashishtMadhavan for insights on training Pong, Shrivastava Anshumali for conversations about randomprojections, and Ozan Sener for discussion of second order methods. We are also grateful to PaulMikesell, Leon Rosenshein, Alex Sergeev and the entire OpusStack Team inside Uber for providingour computing platform and for technical support.

9

Published as a conference paper at ICLR 2018

REFERENCES

Greg Brockman, Vicki Cheung, Ludwig Pettersson, Jonas Schneider, John Schulman, Jie Tang, andWojciech Zaremba. Openai gym, 2016.

Francesco Camastra and Alessandro Vinciarelli. Estimating the intrinsic dimension of data with afractal-based method. IEEE Transactions on pattern analysis and machine intelligence, 24(10):1404–1407, 2002.

Wenlin Chen, James T. Wilson, Stephen Tyree, Kilian Q. Weinberger, and Yixin Chen. Compressingneural networks with the hashing trick. CoRR, abs/1504.04788, 2015. URL http://arxiv.org/abs/1504.04788.

Yann Dauphin and Yoshua Bengio. Big neural networks waste capacity. CoRR, abs/1301.3583,2013. URL http://arxiv.org/abs/1301.3583.

Yann Dauphin, Razvan Pascanu, Caglar Gulcehre, Kyunghyun Cho, Surya Ganguli, and YoshuaBengio. Identifying and attacking the saddle point problem in high-dimensional non-convex op-timization. CoRR, abs/1406.2572, 2014. URL http://arxiv.org/abs/1406.2572.

Misha Denil, Babak Shakibi, Laurent Dinh, Nando de Freitas, et al. Predicting parameters in deeplearning. In Advances in Neural Information Processing Systems, pp. 2148–2156, 2013.

Keinosuke Fukunaga and David R Olsen. An algorithm for finding intrinsic dimensionality of data.IEEE Transactions on Computers, 100(2):176–183, 1971.

Raja Giryes, Guillermo Sapiro, and Alexander M Bronstein. Deep neural networks with randomgaussian weights: a universal classification strategy? IEEE Trans. Signal Processing, 2016.

Xavier Glorot and Yoshua Bengio. Understanding the difficulty of training deep feedforward neuralnetworks. In Artificial Intelligence and Statistics, 2010.

Ian J. Goodfellow, Oriol Vinyals, and Andrew M. Saxe. Qualitatively characterizing neural networkoptimization problems. In International Conference on Learning Representations (ICLR 2015),2015. URL http://arxiv.org/abs/1412.6544.

Song Han, Huizi Mao, and William J Dally. Deep compression: Compressing deep neural networkswith pruning, trained quantization and huffman coding. ICLR, 2016.

Kaiming He, Xiangyu Zhang, Shaoqing Ren, and Jian Sun. Delving deep into rectifiers: Surpassinghuman-level performance on imagenet classification. In CVPR, 2015.

Geoffrey E Hinton and Drew Van Camp. Keeping neural networks simple by minimizing the de-scription length of the weights. In Proceedings of the sixth annual conference on Computationallearning theory, pp. 5–13. ACM, 1993.

Forrest N Iandola, Song Han, Matthew W Moskewicz, Khalid Ashraf, William J Dally, and KurtKeutzer. Squeezenet: Alexnet-level accuracy with 50x fewer parameters and¡ 0.5 mb model size.arXiv preprint arXiv:1602.07360, 2016.

Lukasz Kaiser, Aidan N. Gomez, Noam Shazeer, Ashish Vaswani, Niki Parmar, Llion Jones, andJakob Uszkoreit. One model to learn them all. CoRR, abs/1706.05137, 2017. URL http://arxiv.org/abs/1706.05137.

Balazs Kegl. Intrinsic dimension estimation using packing numbers. In Advances in neural infor-mation processing systems, pp. 697–704, 2003.

Diederik Kingma and Jimmy Ba. Adam: A method for stochastic optimization. arXiv preprintarXiv:1412.6980, 2014.

James Kirkpatrick, Razvan Pascanu, Neil Rabinowitz, Joel Veness, Guillaume Desjardins, Andrei A.Rusu, Kieran Milan, John Quan, Tiago Ramalho, Agnieszka Grabska-Barwinska, Demis Hass-abis, Claudia Clopath, Dharshan Kumaran, and Raia Hadsell. Overcoming catastrophic forget-ting in neural networks. Proceedings of the National Academy of Sciences, 114(13):3521–3526,2017. doi: 10.1073/pnas.1611835114. URL http://www.pnas.org/content/114/13/3521.abstract.

10

Published as a conference paper at ICLR 2018

Alex Krizhevsky and Geoffrey Hinton. Learning multiple layers of features from tiny images. Tech-nical report, 2009.

Quoc Le, Tamas Sarlos, and Alex Smola. Fastfood-approximating kernel expansions in loglineartime. In ICML, 2013.

Elizaveta Levina and Peter J Bickel. Maximum likelihood estimation of intrinsic dimension. InAdvances in neural information processing systems, pp. 777–784, 2005.

Ping Li, T. Hastie, and K. W. Church. Very sparse random projections. In KDD, 2006.

C. Louizos, K. Ullrich, and M. Welling. Bayesian Compression for Deep Learning. ArXiv e-prints,May 2017.

V. Mnih, K. Kavukcuoglu, D. Silver, A. Graves, I. Antonoglou, D. Wierstra, and M. Riedmiller.Playing Atari with Deep Reinforcement Learning. ArXiv e-prints, December 2013.

Shaoqing Ren, Kaiming He, Ross Girshick, and Jian Sun. Faster r-cnn: Towards real-time objectdetection with region proposal networks. In Advances in neural information processing systems,pp. 91–99, 2015.

Jorma Rissanen. Modeling by shortest data description. Automatica, 14(5):465–471, 1978.

Olga Russakovsky, Jia Deng, Hao Su, Jonathan Krause, Sanjeev Satheesh, Sean Ma, ZhihengHuang, Andrej Karpathy, Aditya Khosla, Michael Bernstein, et al. Imagenet large scale visualrecognition challenge. International Journal of Computer Vision, 2015.

T. Salimans, J. Ho, X. Chen, and I. Sutskever. Evolution Strategies as a Scalable Alternative toReinforcement Learning. ArXiv e-prints, March 2017.

Kenneth Stanley and Risto Miikkulainen. Evolving neural networks through augmenting topologies.Evolutionary computation, 10(2):99–127, 2002.

Joshua B Tenenbaum, Vin De Silva, and John C Langford. A global geometric framework fornonlinear dimensionality reduction. science, 290(5500):2319–2323, 2000.

Tijmen Tieleman and Geoffrey Hinton. Lecture 6.5-rmsprop: Divide the gradient by a runningaverage of its recent magnitude. COURSERA: Neural networks for machine learning, 4(2):26–31, 2012.

Emanuel Todorov, Tom Erez, and Yuval Tassa. Mujoco: A physics engine for model-based control.In Intelligent Robots and Systems (IROS), 2012 IEEE/RSJ International Conference on, pp. 5026–5033. IEEE, 2012.

Wei Wen, Chunpeng Wu, Yandan Wang, Yiran Chen, and Hai Li. Learning structured sparsity indeep neural networks. In NIPS, 2016.

Zichao Yang, Marcin Moczulski, Misha Denil, Nando de Freitas, Alex Smola, Le Song, and ZiyuWang. Deep fried convnets. In Proceedings of the IEEE International Conference on ComputerVision, pp. 1476–1483, 2015.

Chiyuan Zhang, Samy Bengio, Moritz Hardt, Benjamin Recht, and Oriol Vinyals. Understandingdeep learning requires rethinking generalization. ICLR, 2017.

11

Published as a conference paper at ICLR 2018

SUPPLEMENTARY INFORMATION FOR:MEASURING THE INTRINSIC DIMENSION OF

OBJECTIVE LANDSCAPES

S5 ADDITIONAL MNIST RESULTS AND INSIGHTS

S5.1 SWEEPING DEPTHS AND WIDTHS; MULTIPLE RUNS TO ESTIMATE VARIANCE

In the main paper, we attempted to find dint90 across 20 FC networks with various depths andwidths. A grid sweep of number of hidden layers from {1,2,3,4,5} and width of each hidden layerfrom {50,100,200,400} is performed, and all 20 plots are shown in Fig. S6. For each d we take 3runs and plot the mean and variance with blue dots and blue error bars. dint90 is indicated in plots(darkened blue dots) by the dimension at which the median of the 3 runs passes 90% performancethreshold. The variance of dint90 is estimated using 50 bootstrap samples. Note that the variance ofboth accuracy and measured dint90 for a given hyper-parameter setting are generally small, and themean of performance monotonically increases (very similar to the single-run result) as d increases.This illustrates that the difference between lucky vs. unlucky random projections have little impacton the quality of solutions, while the subspace dimensionality has a great impact. We hypothesizethat the variance due to different P matrices will be smaller than the variance due to different randominitial parameter vectors θ(D)

0 because there are dD i.i.d. samples used to create P (at least in thedense case) but onlyD samples used to create θ(D)

0 , and aspects of the network depending on smallernumbers of random samples will exhibit greater variance. Hence, in some other experiments we relyon single runs to estimate the intrinsic dimension, though slightly more accurate estimates could beobtained via multiple runs.

In similar manner to the above, in Fig. S7 we show the relationship between dint90 and D across 20networks but using a per-model, directly trained baseline. Most baselines are slightly below 100%accuracy. This is in contrast to Fig. 3, which used a simpler global baseline of 100% across allmodels. Results are qualitatively similar but with slightly lower intrinsic dimension due to slightlylower thresholds.

S5.2 ADDITIONAL DETAILS ON SHUFFLED MNIST DATASETS

Two kinds of shuffled MNIST datasets are considered:

• The shuffled pixel dataset: the label for each example remains the same as the normaldataset, but a random permutation of pixels is chosen once and then applied to all images inthe training and test sets. FC networks solve the shuffled pixel datasets exactly as easily asthe base dataset, because there is no privileged ordering of input dimension in FC networks;all orderings are equivalent.

• The shuffled label dataset: the images remain the same as the normal dataset, but labelsare randomly shuffled for the entire training set. Here, as in (Zhang et al., 2017), we onlyevaluate training accuracy, as test set accuracy remains forever at chance level (the trainingset X and y convey no information about test set p(y|X), because the shuffled relationshipin test is independent of that of training).

On the full shuffled label MNIST dataset (50k images), we trained an FC network (L = 5,W = 400,which had dint90 = 750 on standard MNIST), it yields dint90 = 190k. We can interpret this asrequiring 3.8 floats to memorize each random label (at 90% accuracy). Wondering how this scaleswith dataset size, we estimated dint90 on shuffled label versions of MNIST at different scales andfound curious results, shown in Table S2 and Fig. S8. As the dataset memorized becomes smaller,the number of floats required to memorize each label becomes larger. Put another way, as dataset size

12

Published as a conference paper at ICLR 2018

0 500 1000 1500Subspace dim d

0.2

0.4

0.6

0.8

1.0Va

lidat

ion

accu

racy

width 50, depth 1

0 500 1000 1500Subspace dim d

0.2

0.4

0.6

0.8

1.0width 50, depth 2

0 500 1000 1500Subspace dim d

0.2

0.4

0.6

0.8

1.0width 50, depth 3

0 500 1000 1500Subspace dim d

0.2

0.4

0.6

0.8

1.0width 50, depth 4

0 500 1000 1500Subspace dim d

0.2

0.4

0.6

0.8

1.0width 50, depth 5

0 500 1000 1500Subspace dim d

0.2

0.4

0.6

0.8

1.0

Valid

atio

n ac

cura

cy

width 100, depth 1

0 500 1000 1500Subspace dim d

0.2

0.4

0.6

0.8

1.0width 100, depth 2

0 500 1000 1500Subspace dim d

0.2

0.4

0.6

0.8

1.0width 100, depth 3

0 500 1000 1500Subspace dim d

0.2

0.4

0.6

0.8

1.0width 100, depth 4

0 500 1000 1500Subspace dim d

0.2

0.4

0.6

0.8

1.0width 100, depth 5

0 500 1000 1500Subspace dim d

0.2

0.4

0.6

0.8

1.0

Valid

atio

n ac

cura

cy

width 200, depth 1

0 500 1000 1500Subspace dim d

0.2

0.4

0.6

0.8

1.0width 200, depth 2

0 500 1000 1500Subspace dim d

0.2

0.4

0.6

0.8

1.0width 200, depth 3

0 500 1000 1500Subspace dim d

0.2

0.4

0.6

0.8

1.0width 200, depth 4

0 500 1000 1500Subspace dim d

0.2

0.4

0.6

0.8

1.0width 200, depth 5

0 500 1000 1500Subspace dim d

0.2

0.4

0.6

0.8

1.0

Valid

atio

n ac

cura

cy

width 400, depth 1

0 500 1000 1500Subspace dim d

0.2

0.4

0.6

0.8

1.0width 400, depth 2

0 500 1000 1500Subspace dim d

0.2

0.4

0.6

0.8

1.0width 400, depth 3

0 500 1000 1500Subspace dim d

0.2

0.4

0.6

0.8

1.0width 400, depth 4

0 500 1000 1500Subspace dim d

0.2

0.4

0.6

0.8

1.0width 400, depth 5

Figure S6: A sweep of FC networks on MNIST. Each column contains networks of the same depth,and each row those of the same number of hidden nodes in each of its layers. Mean and variance ateach d is shown by blue dots and blue bars. dint90 is found by dark blue dots, and the variance of itis indicated by red bars spanning in the d axis.

105 106

Number of parameters D400

500

600

700

800

900

1000

Intri

nsic

dim

ensio

n d i

nt90

layer width50100200400

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0De

pth

Figure S7: Measured intrinsic dimension dint90 vs number of parametersD for 20 models of varyingwidth (from 50 to 400) and depth (number of hidden layers from 1 to 5) trained on MNIST. Thevertical red interval is the standard derivation of measured dint90. As opposed to Fig. 3, which useda global, shared baseline across all models, here a per-model baseline is used. The number of nativeparameters varies by a factor of 24.1, but dint90 varies by only 1.42. The per-model baseline resultsin higher measured dint90 for larger models because they have a higher baseline performance thanthe shallower models.

13

Published as a conference paper at ICLR 2018

Table S2: dint90 required to memorize shuffled MNIST labels. As dataset size grows, memorizationbecomes more efficient, suggesting a form of “generalization” from one part of the training set toanother, even though labels are random.

Fraction of MNIST training set dint90 Floats per label100% 190k 3.8

50% 130k 5.210% 90k 18.0

0 50000 100000 150000 200000 250000 300000Intrinsic dim d

0.0

0.2

0.4

0.6

0.8

1.0

Tra

inin

gac

cura

cy

100%

50%

10%

baseline

Figure S8: Training accuracy vs. subspace dimension d for a FC networks (W=400, L=5) trainedon a shuffled label version of MNIST containing 100%, 50%, and 10% of the dataset.

increases, the intrinsic dimension also increases, but not as fast as linearly. The best interpretationis not yet clear, but one possible interpretation is that networks required to memorize large trainingsets make use of shared machinery for memorization. In other words, though performance doesnot generalize to a validation set, generalization within a training set is non-negligible even thoughlabels are random.

S5.3 TRAINING STABILITY

An interesting tangential observation is that random subspace training can in some cases make op-timization more stable. First, it helps in the case of deeper networks. Fig. S9 shows training resultsfor FC networks with up to 10 layers. SGD with step 0.1, and ReLUs with He initialization is used.Multiple networks failed at depths 4, and all failed at depths higher than 4, despite the activationfunction and initialization designed to make learning stable (He et al., 2015). Second, for MNISTwith shuffled labels, we noticed that it is difficult to reach high training accuracy using the directtraining method with SGD, though both subspace training with SGD and either type of training withAdam reliably reach 100% memorization as d increases (see Fig. S8).

Because each random basis vector projects across all D direct parameters, the optimization problemmay be far better conditioned in the subspace case than in the direct case. A related potentialdownside is that projecting acrossD parameters which may have widely varying scale could result inignoring parameter dimensions with tiny gradients. This situation is similar to that faced by methodslike SGD, but ameliorated by RMSProp, Adam, and other methods that rescale per-dimension stepsizes to account for individual parameter scales. Though convergence of the subspace approachseems robust, further work may be needed to improve network amenability to subspace training: forexample by ensuring direct parameters are similarly scaled by clever initialization or by inserting apre-scaling layer between the projected subspace and the direct parameters themselves.

S5.4 THE ROLE OF OPTIMIZERS

Another finding through our experiments with MNIST FC networks has to do with the role of opti-mizers. The same set of experiments are run with both SGD (learning rate 0.1) and ADAM (learning

14

Published as a conference paper at ICLR 2018

0 2 4 6 8 10Depth

0.0

0.2

0.4

0.6

0.8

1.0

Test

ing

Accu

racy

directd = 50d = 200d = 500d = 1000

0 2 4 6 8 10Depth

0.0

0.5

1.0

1.5

2.0

2.5

Test

ing

NLL

directd = 50d = 200d = 500d = 1000

(a) Accuracy (b) Negative Log Likelihood (NLL)

Figure S9: Results of subspace training versus the number of layers in a fully connected networktrained on MNIST. The direct method always fail to converge when L > 5, while subspace trainingyields stable performance across all depths.

rate 0.001), allowing us to investigate the impact of stochastic optimizers on the intrinsic dimensionachieved.

The intrinsic dimension dint90 are reported in Fig. S10 (a)(b). In addition to two optimizers we alsouse two baselines: Global baseline that is set up as 90% of best performance achieved across allmodels, and individual baseline that is with regards to the performance of the same model in directtraining.

S6 ADDITIONAL REINFORCEMENT LEARNING RESULTS AND DETAILS

S6.1 DQN EXPERIMENTS

DQN on Cartpole We start with a simple classic control game CartPole−v0 in OpenAIGym (Brockman et al., 2016). A pendulum starts upright, and the goal is to prevent it from fallingover. The system is controlled by applying a force of LEFT or RIGHT to the cart. The full gameends when one of two failure conditions is satisfied: the cart moves more than 2.4 units from thecenter (where it started), or the pole is more than 15 degrees from vertical (where it started). Areward of +1 is provided for every time step as long as the game is going. We further created twoeasier environments Pole and Cart, each confined by one of the failure modes only.

A DQN is used, where the value network is parameterized by an FC (L = 2,W = 400). For eachsubspace d at least 5 runs are conducted, the mean of which is used to computed dint90, and thebaseline is set as 195.07. The results are shown in Fig. S11. The solid line connects mean rewardswithin a run over the last 100 episodes, across different ds. Due to the noise-sensitiveness of RLgames the course is not monotonic any more. The intrinsic dimension for CartPole, Pole and Cartis dint90 = 25, 23 and 7, respectively. This reveals that the difficulty of optimization landscape ofthese games is remarkably low, as well as interesting insights such as driving a cart is much easierthan keeping a pole straight, the latter being the major cause of difficulty when trying to do both.

S6.2 EVOLUTIONARY STRATEGIES (ES) COMPLETE RESULTS

We carry out with ES 3 RL tasks: InvertedPendulum−v1, Humanoid−v1, Pong−v0. The hy-perparameter settings for training are in Table S3.

Inverted pendulum The InvertedPendulum−v1 environment uses the MuJoCo physics simu-lator (Todorov et al., 2012) to instantiate the same problem as CartPole−v0 in a realistic setting.We expect that even with richer environment dynamics, as well as a different RL algorithm – ES –the intrinsic dimensionality should be similar. As seen in Fig. 5, the measured intrinsic dimensional-ity dint90 = 4 is of the same order of magnitude, but smaller. Interestingly, although the environment

7CartPole−v0 is considered as “solved” when the average reward over the last 100 episodes is 195

15

Published as a conference paper at ICLR 2018

105 106

# Model parameters400

500

600

700

800

900

1000

Intri

nsic

dim

din

t

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

Dept

h

(a) Global baseline

105 106

# Model parameters400

500

600

700

800

900

1000

Intri

nsic

dim

din

t

layer width

50

100

200

400

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

Dept

h

(b) Individual baseline

Figure S10: The role of optimizers on MNIST FC networks. The transparent dots indicate SGDresults, and opaque dots indicate Adam results. Adam generally yields higher intrinsic dimensionsbecause higher baselines are achieved, especially when individual baselines in (b) are used. Notethat the Adam points are slightly different between Fig. 3 and Fig. S7, because in the former weaverage over three runs, and in the latter we show one run each for all optimization methods.

100 101

Subspace dim d

0

50

100

150

200

Ave

rage

Rew

ard

over

last

100

epi.

baseline

90% baseline

100 101

Subspace dim d

0

50

100

150

200

Ave

rage

Rew

ard

over

last

100

epi.

baseline

90% baseline

100 101

Subspace dim d

50

100

150

200

Ave

rage

Rew

ard

over

last

100

epi.

baseline

90% baseline

(a) CartPole (dint90 = 25) (b) Pole (dint90 = 23) (c) Cart (dint90 = 7)

Figure S11: Subspace training of DQN on CartPole game. Shown as dots are rewards collectedthrough a game run averaged over the last 100 episodes, under each subspace training of DQN, andeach game environment. The line connects mean rewards across different ds.

dynamics are more complex than in CartPole−v0, using ES rather than DQN seems to induce asimpler objective landscape.

Learning to walk A more challenging problem is Humanoid−v1 in MuJoCo simulator. Intu-itively, one might believe that learning to walk is a more complex task than classifying images. Ourresults show the contrary – that the learned intrinsic dimensionality of dint90 = 700 is similar tothat of MNIST on a fully-connected network (dint90 = 650) but significantly less than that of evena convnet trained on CIFAR-10 (dint90 = 2, 500). Fig. 5 shows the full results. Interestingly, we

16

Published as a conference paper at ICLR 2018

`2 penalty Adam LR ES σ IterationsInvertedPendulum−v1 1× 10−8 3× 10−1 2× 10−2 1000

Humanoid−v1 5× 10−3 3× 10−2 2× 10−2 2000Pong−v0 5× 10−3 3× 10−2 2× 10−2 500

Table S3: Hyperparameters used in training RL tasks using ES. σ refers to the parameter perturbationnoise used in ES. Default Adam parameters of β1 = 0.9, β2 = 0.999, ε = 1× 10−7 were used.

begin to see training runs reach the threshold as early as d = 400, with the median performancesteadily increasing with d.

Atari Pong Finally, using a base convnet of approximately D = 1M in the Pong−v0 pixels-to-actions environment (using 4-frame stacking). The agent receives an image frame (size of 210 ×160 × 3) and the action is to move the paddle UP or DOWN. We were able to determine dint90 =6, 000.

S7 THREE METHODS OF RANDOM PROJECTION

Scaling the random subspace training procedure to large problems requires an efficient way to mapfrom Rd into a random d-dimensional subspace of RD that does not necessarily include the origin.Algebraically, we need to left-multiply a vector of parameters v ∈ Rd by a random matrix M ∈RD×d, whose columns are orthonormal, then add an offset vector θ0 ∈ RD. If the low-dimensionalparameter vector in Rd is initialized to zero, then specifying an offset vector is equivalent to choosingan initialization point in the original model parameter space RD.

A naıve approach to generating the random matrix M is to use a dense D× d matrix of independentstandard normal entries, then scale each column to be of length 1. The columns will be approxi-mately orthogonal if D is large because of the independence of the entries. Although this approachis sufficient for low-rank training of models with few parameters, we quickly run into scaling limitsbecause both matrix-vector multiply time and storage of the matrix scale according to O(Dd). Wewere able to successfully determine the intrinsic dimensionality of MNIST (d=225) using a LeNet(D=44,426), but were unable to increase d beyond 1,000 when applying a LeNet (D=62,006) toCIFAR-10, which did not meet the performance criterion to be considered the problems intrinsicdimensionality.

Random matrices need not be dense for their columns to be approximately orthonormal. In fact, amethod exists for “very sparse” random projections (Li et al., 2006), which achieves a density of 1√

D.

To construct the D × d matrix, each entry is chosen to be nonzero with probability 1√D

. If chosen,then with equal probability, the entry is either positive or negative with the same magnitude in eithercase. The density of 1√

Dimplies

√Dd nonzero entries, or O(

√Dd) time and space complexity.

Implementing this procedure allowed us to find the intrinsic dimension of d=2,500 for CIFAR-10using a LeNet mentioned above. Unfortunately, when using Tensorflow’s SparseTensor imple-mentation we did not achieve the theoretical

√D-factor improvement in time complexity (closer to

a constant 10x). Nonzero elements also have a significant memory footprint of 24 bytes, so we couldnot scale to larger problems with millions of model parameters and large intrinsic dimensionalities.

We need not explicitly form and store the transformation matrix. The Fastfood transform (Le et al.,2013) was initially developed as an efficient way to compute a nonlinear, high-dimensional featuremap φ(x) for a vector x. A portion of the procedure involves implicitly generating a D × d matrixwith approximately uncorrelated standard normal entries, using only O(D) space, which can bemultiplied by v in O(D log d) time using a specialized method. The method relies on the fact thatHadamard matrices multiplied by Gaussian vectors behave like dense Gaussian matrices. In detail,to implicitly multiply v by a random square Gaussian matrix M with side-lengths equal to a powerof two, the matrix is factorized into multiple simple matrices: M = HGΠHB, where B is arandom diagonal matrix with entries +-1 with equal probability, H is a Hadamard matrix, Π is arandom permutation matrix, and G is a random diagonal matrix with independent standard normalentries. Multiplication by a Hadamard matrix can be done via the Fast Walsh-Hadamard Transform

17

Published as a conference paper at ICLR 2018

in O(d log d) time and takes no additional space. The other matrices have linear time and spacecomplexities. When D > d, multiple independent samples of M can be stacked to increase theoutput dimensionality. When d is not a power of two, we can zero-pad v appropriately. StackingDd samples of M results in an overall time complexity of O(D

d d log d) = O(D log d), and a spacecomplexity of O(D

d d) = O(D). In practice, the reduction in space footprint allowed us to scale tomuch larger problems, including the Pong RL task using a 1M parameter convolutional network forthe policy function.

Table S4 summarizes the performance of each of the three methods theoretically and empirically.

Time complexity Space complexity D = 100k D = 1M D = 60MDense O(Dd) O(Dd) 0.0169 s 1.0742 s* 4399.1 s*Sparse O(

√Dd) O(

√Dd) 0.0002 s 0.0019 s 0.5307 s*

Fastfood O(D log d) O(D) 0.0181 s 0.0195 s 0.7949 s

Table S4: Comparison of theoretical complexity and average duration of a forward+backward passthrough M (in seconds). d was fixed to 1% of D in each measurement. D = 100k is approximatelythe size of an MNIST fully-connected network, and D = 60M is approximately the size of AlexNet.The Fastfood timings are based on a Tensorflow implementation of the Fast Walsh-Hadamard Trans-form, and could be drastically reduced with an efficient CUDA implementation. Asterisks mean thatwe encountered an out-of-memory error, and the values are extrapolated from the largest successfulrun (a few powers of two smaller). For example, we expect sparse to outperform Fastfood if it didn’trun into memory issues.

Figure S4 compares the computational time for direct and subspace training (various projections)methods for each update. Our subspace training is more computational expensive, because thesubspace training method has to propagate the signals through two modules: the layers of neuralnetworks, and the projection between two spaces. The direct training only propagates signals in thelayers of neural networks. We have made efforts to reduce the extra computational cost. For exam-ple, the sparse projection less than doubles the time cost for a large range of subspace dimensions.

102 103 104 105

Intrinsic dim d0.000

0.005

0.010

0.015

0.020

0.025

0.030

time

(s)

densesparsefastfooddirect

102 103 104 105 106

Intrinsic dim d0.000

0.005

0.010

0.015

0.020

time

(s) dense

sparsefastfooddirect

Figure S12: MNIST compute time for direct vs. various projection methods for 100k parameters(left) and 1M parameters (right).

S8 ADDITIONAL CIFAR-10 RESULTS

FC networks We consider the CIFAR-10 dataset and test the same set of FC and LeNet architec-tures as on MNIST. For FC networks, dint90 values for all 20 networks are shown in Fig. S13 (a)plotted against the native dimension D of each network; D changes by a factor of 12.16 betweenthe smallest and largest networks, but dint90 changes over this range by a factor of 5.0. However,much of this change is due to change of baseline performance. In Fig. S13 (b), we instead computethe intrinsic dimension with respect to a global baseline: 50% validation accuracy. dint90 changesover this range by a factor of 1.52. This indicates that various FC networks share similar intrin-sic dimension (dint90 = 5000 ∼ 8000) to achieve the same level of task performance. For LeNet(D = 62, 006), the validation accuracy vs. subspace dimension d is shown in Fig. S14 (b), the cor-responding dint90 = 2900. It yields a compression rate of 5%, which is 10 times larger than LeNeton MNIST. It shows that CIFAR-10 images are significantly more difficult to be correctly classified

18

Published as a conference paper at ICLR 2018

106

Number of parameters D0

2500

5000

7500

10000

12500

15000

Intri

nsic

dim

din

t

layer width

50

100

200

400

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

Dept

h

(a) Individual baseline

106

Number of parameters D0

2500

5000

7500

10000

12500

15000

Intri

nsic

dim

din

t

layer width

50

100

200

400

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

Dept

h

(b) Global baseline

Figure S13: Intrinsic dimension of FC networks with various width and depth on the CIFAR-10dataset. In (b), we use a simple 50% baseline globally.

than MNIST. In another word, CIFAR-10 is a harder problem than MNIST, especially given the factthat the notion of “problem solved” (baseline performance) is defined as 99% accuracy on MNISTand 58% accuracy on CIFAR-10. On the CIFAR-10 dataset, as d increases, subspace training tendsto overfitting; we study the role of subspace training as a regularizer below.

ResNet vs. LeNet We test ResNets, compare to LeNet, and find they make efficient use of pa-rameters. We adopt the smallest 20-layer structure of ResNet with 280k parameters, and find outin Fig. S14 (b) that it reaches LeNet baseline with dint90 = 1000 ∼ 2000 (lower than the dint90 ofLeNet), while takes a larger dint90 (20, 000 ∼ 50, 000) to reach reach its own, much higher baseline.

The role of regularizers Our subspace training can be considered as a regularization scheme, as itrestricts the solution set. We study and compare its effects with two traditional regularizers with anFC network (L=2, W=200) on CIFAR-10 dataset, including `2 penalty on the weights (i.e., weightdecay) and Dropout.

• `2 penalty Various amount of `2 penalty from {10−2, 10−3, 5× 10−4, 10−4, 10−5, 0}are considered. The accuracy and negative log-likelihood (NLL) are reported in Fig. S15(a) (b), respectively. As expected, larger amount of weight decay reduces the gap be-tween training and testing performance for both direct and subspace training methods, andeventually closes the gap (i.e., `2 penalty = 0.01). Subspace training itself exhibits strongregularization ability, especially when d is small, at which the performance gap betweentraining and testing is smaller.

• Dropout Various dropout rates from {0.5, 0.4, 0.3, 0.2, 0.1, 0} are considered. The ac-curacy and NLL are reported in Fig. S16. Larger dropout rates reduce the gap betweentraining and testing performance for both direct and subspace training methods. Whenobserving testing NLL, subspace training tends to overfit the training dataset less.

• Subspace training as implicit regularization Subspace training method performs implicitregularization, as it restricts the solution set. We visualized the testing NLL in Fig. S17.Subspace training method outperforms direct method when d is properly chosen (when `2penalty< 5×10−4, or dropout rate< 0.1 ), suggesting the potential of this method as a betteralternative to traditional regularizers. When d is large, the method also overfits the training

19

Published as a conference paper at ICLR 2018

0 5000 10000 15000 20000 25000 30000Subspace dim d

0.35

0.40

0.45

0.50

Val

idat

ion

accu

racy

baseline

90% baseline

(a) FC (W = 200, L = 2)

0 5000 10000 15000 20000 25000 30000Subspace dim d

0.3

0.4

0.5

0.6

Val

idat

ion

accu

racy

baseline

90% baseline

(b) LeNet

0 10000 20000 30000 40000 50000Subspace dim d

0.4

0.5

0.6

0.7

0.8

Val

idat

ion

accu

racy

ResNet baseline

90% ResNet baseline

90% LeNet baseline

(c) ResNet

Figure S14: Validation accuracy of an FC network, LeNet and ResNet on CIFAR with different sub-space dimension d. In (a)(b), the variance of validation accuracy and measured dint90 are visualizedas the blue vertical and red horizontal error bars, respectively. Subspace method surpasses the 90%baseline on LeNet at d between 1000 and 2000, 90% of ResNet baseline between 20k and 50k.

dataset. Note that the these methods perform regularization in different ways: weight decayenforces the learned weights concentrating around zeros, while subspace training directlyreduces the number of dimensions of the solution space.

S9 IMAGENET

To investigate even larger problems, we attempted to measure dint90 for an ImageNet classificationnetwork. We use a relatively smaller network, SqueezeNet by Iandola et al. (2016), with 1.24M pa-rameters. Larger networks suffered from memory issues. A direct training produces Top-1 accuracyof 55.5%. We vary intrinsic dimension from 50k, 100k, 200k, 500k, 800k, and record the validationaccuracies as shown in Fig. S18. The training of each intrinsic dimension takes about 6 to 7 days,distributed across 4 GPUs. Due to limited time, training on ImageNet has not yet produced a reliableestimate for dint90 except that it is over 500k.

S10 INVESTIGATION OF CONVOLUTIONAL NETWORKS

Since the learned dint90 can be used as a robust measure to study the fitness of neural net-work architectures for specific tasks, we further apply it to understand the contribution of eachcomponent in convolutional networks for image classification task. The convolutional network

20

Published as a conference paper at ICLR 2018

0 20000 40000Subspace dimension d

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

Accu

racy

2 penalty = 0.01TestingTrainingTesting: directTraining: direct

0 20000 40000Subspace dimension d

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.02 penalty = 0.001

0 20000 40000Subspace dimension d

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.02 penalty = 0.0005

0 20000 40000Subspace dimension d

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.02 penalty = 0.0001

0 20000 40000Subspace dimension d

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.02 penalty = 1e-05

0 20000 40000Subspace dimension d

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.02 penalty = 0

(a) Accuracy

0 20000 40000Subspace dim d

123456789

NLL

2 penalty = 0.01

TestingTrainingTesting: directTraining: direct

0 20000 40000Subspace dim d

0

1

2

3

4

5

2 penalty = 0.001

0 20000 40000Subspace dim d

0

1

2

3

4

5

2 penalty = 0.0005

0 20000 40000Subspace dim d

0

1

2

3

4

5

2 penalty = 0.0001

0 20000 40000Subspace dim d

0

1

2

3

4

5

2 penalty = 1e-05

0 20000 40000Subspace dim d

0

1

2

3

4

5

2 penalty = 0

(b) NLL

Figure S15: Comparing regularization induced by `2 penalty and subspace training. Weight decayinteracts with dint90 since it changes the objective landscapes through various loss functions.

0 20000 40000Subspace dimension d

0.160.240.320.400.480.560.640.720.800.88

Accu

racy

Dropout rate = 0.5TestingTrainingTesting: directTraining: direct

0 20000 40000Subspace dimension d

0.160.240.320.400.480.560.640.720.800.88

Dropout rate = 0.4

0 20000 40000Subspace dimension d

0.160.240.320.400.480.560.640.720.800.88

Dropout rate = 0.3

0 20000 40000Subspace dimension d

0.160.240.320.400.480.560.640.720.800.88

Dropout rate = 0.2

0 20000 40000Subspace dimension d

0.160.240.320.400.480.560.640.720.800.88

Dropout rate = 0.1

0 20000 40000Subspace dimension d

0.160.240.320.400.480.560.640.720.800.88

Dropout rate = 0.0

(a) Accuracy

0 20000 40000Subspace dim d

1

2

3

NLL

Dropout rate = 0.5TestingTrainingTesting: directTraining: direct

0 20000 40000Subspace dim d

1

2

3

Dropout rate = 0.4

0 20000 40000Subspace dim d

1

2

3

Dropout rate = 0.3

0 20000 40000Subspace dim d

1

2

3

Dropout rate = 0.2

0 20000 40000Subspace dim d

1

2

3

Dropout rate = 0.1

0 20000 40000Subspace dim d

1

2

3

Dropout rate = 0.0

(b) NLL

Figure S16: Comparing regularization induced by Dropout and subspace training. Dropout interactswith dint90 since it changes the objective landscapes through randomly removing hidden units of theextrinsic neural networks.

2 penalty

0.01

0.001

0.0001

1e-05

Intrinsic dim d

010000

2000030000

4000050000

Valid

atio

n NL

L

1.202.062.933.794.655.516.387.248.108.96

10000 20000 30000 40000

Dropout rate

0

0.1

0.2

0.3

0.4

0.5

Intrinsic dim d

010000

2000030000

4000050000

Valid

atio

n NL

L

1.201.361.511.671.821.982.132.292.442.60

10000 20000 30000 40000

(a) `2 penalty (b) Dropout

Figure S17: Comparing regularization induced by `2 penalty, Dropout and subspace training. Thegray surface and black line indicate direct training.

21

Published as a conference paper at ICLR 2018

100000 200000 300000 400000 500000 600000 700000 800000Subspace dim d

0.2

0.3

0.4

0.5

Val

idat

ion

accu

racy

0.2933

0.36090.3895

Baseline

90% baseline

Figure S18: Validation accuracy of SqueezeNet on ImageNet with different d. At d = 500k theaccuracy reaches 34.34%, which is not yet past the threshold required to estimate dint90.

is a special case of FC network in two aspects: local receptive fields and weight-tying. Lo-cal receptive fields force each filter to “look” only at a small, localized region of the im-age or layer below. Weight-tying enforces that each filter shares the same weights, whichreduces the number of learnable parameters. We performed control experiments to investi-gate the degree to which each component contributes. Four variants of LeNet are considered:

• Standard LeNet 6 kernels (5×5) – max-pooling (2×2) – 16 kernels (5×5) – max-pooling(2× 2) – 120 FC – 84 FC – 10 FC• Untied-LeNet The same architecture with the standard LeNet is employed, except that

weights are unshared, i.e., a different set of filters is applied at each different patch ofthe input. For example in Keras, the LocallyConnected2D layer is used to replace theConv2D layer.

• FCTied-LeNet The same set of filters is applied at each different patch of the input. webreak local connections by applying filters to global patches of the input. Assume the imagesize is H ×H , the architecture is 6 kernels ((2H−1) × (2H−1)) – max-pooling (2 × 2)– 16 kernels ((H−1) × (H−1)) – max-pooling (2 × 2) – 120 FC – 84 FC – 10 FC. Thepadding type is same.

• FC-LeNet Neither local connections or tied weights is employed, we mimic LeNet withits FC implementation. The same number of hidden units as the standard LeNet are used ateach layer.

The results are shown in Fig. S19 (a)(b). We set a crossing-line accuracy (i.e., threshold) for eachtask, and investigate dint90 needed to achieve it. For MNIST and CIFAR-10, the threshold is 90%and 45%, respectively. For the above LeNet variants, dint90 = 290, 600, 425, 2000 on MNIST, anddint90 = 1000, 2750, 2500, 35000 on CIFAR-10. Experiments show both tied-weights and localconnections are important to the model. That tied-weights should matter seems sensible. However,models with maximal convolutions (convolutions covering the whole image) may have had the sameintrinsic dimension as smaller convolutions, but this turns out not to be the case.

S11 SUMMARIZATION OF dint90

We summarize dint90 of the objective landscape on all different problems and neural network archi-tectures in Table S5 and Fig. S20. “SP” indicates shuffled pixel, and “SL” for shuffled label, and“FC-5” for a 5-layer FC. dint90 indicates the minimum number of dimensions of trainable parametersrequired to properly solve the problem, and thus reflects the difficulty level of problems.

22

Published as a conference paper at ICLR 2018

102 103 104

Intrinsic dim d

0.4

0.6

0.8

1.0

Val

idat

ion

accu

racy

Standard LeNet

Untied-LeNet

FCTied-LeNet

FC-LeNet

threshold

103 104

Intrinsic dim d

0.2

0.4

0.6

Val

idat

ion

accu

racy

Standard LeNet

Untied-LeNet

FCTied-LeNet

FC-LeNet

threshold

(a) MNIST (b) CIFAR-10

Figure S19: Validation accuracy of LeNet variants with different subspace dimension d. The con-clusion is that convnets are more efficient than FC nets both due to local connectivity and due toweight tying.

Table S5: Intrinsic dimension of different objective landscapes, determined by dataset and network.

Dataset Network D dint90MNIST FC 199210 750MNIST LeNet 44426 275

CIFAR-10 FC 1055610 9000CIFAR-10 LeNet 62006 2900MNIST-SP FC 199210 750MNIST-SP LeNet 44426 650

MNIST-SL-100% FC-5 959610 190000MNIST-SL-50% FC-5 959610 130000MNIST-SL-10% FC-5 959610 90000

ImageNet SqueezeNet 1248424 >500000CartPole FC 199210 25

Pole FC 199210 23Cart FC 199210 7

Inverted Pendulum FC 562 4Humanoid FC 166673 700Atari Pong ConvNet 1005974 6000

101 102 103 104 105

Subspace dim d

103

104

105

106

107

Num

ber

ofpa

ram

eter

sD

MNIS

T-FC

MNIS

T-LeN

et

Cifar-F

C

Cifar-L

eNet

MNIS

T-SP-F

C

MNIS

T-SP-L

eNet

MNIS

T-SL-1

00%

-FC-5

MNIS

T-SL-5

0%-F

C-5

MNIS

T-SL-1

0%-F

C-5

CartP

ole-F

C

Pole-F

CCar

t-FC

Inve

rtedPen

dulum-F

C

Human

oid-F

C

Atari

Pong-

ConvN

et

Imag

eNet

-Squee

zeNet

Figure S20: Intrinsic dimension of the objective landscapes created by all combinations of datasetand network we tried in this paper.

23